Thermal Model of Eurodish Solar Stirling Engine

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

THERMAL MODEL OF THE EURODISH SOLAR STIRLING

ENGINE

Francisco J. García Granados, Manuel A. Silva Pérez, V. Ruiz-Hernández,

Grupo de Termodinámica y Energías Renovables.


Escuela Técnica Superior de Ingenieros, Universidad de Sevilla.

Camino de Los Descubrimientos s/n, 41092 Sevilla, Spain


Tel: + 34 - 954487232; Fax: +34 - 954487233; e-mail: silva@esi.us.es.

Abstract
One parabolic dish –Stirling engine system has been in operation at the Engineering School of
Seville since March 2004. This system is one of the several Country Reference Units of the
EnviroDish project, and is based on the Eurodish system.

The system has achieved a maximum thermal efficiency (solar to electricity) close to 20%
during operation. The analysis of the different parameters suggests a high potential for
improvement. A thermal model of the main components of the engine package (cavity,
receiver and Stirling engine) can help to evaluate possible modifications of the system and
identify the most promising ones. The development of such thermal model and its comparison
with experimental data gathered during this period, are reported in this work. Model results
exhibit good qualitative agreement with the available measurements. However, the validation
of the model will require measuring more parameters at the cavity, receiver and engine.

Keywords: Solar concentrating systems, parabolic dish, Stirling engine, thermal losses,
thermal model.

------------------------------------

Introduction
The 10 kW parabolic dish - Stirling system installed at the premises of the Sevilla
Engineering School (ESI) is one of the five Country Reference Units of the EnviroDish
project [1], consisting in the installation and operation of several such units, based on the
Eurodish system, in different countries aimed to gathering reliable O&M data that help to
improve and fine-tune the technology. The Centre for New Energy Technologies (CENTER),
as local partner of the project, financed part of the system –with subsidies from the
Andalousian Regional Government and support from Endesa Distribución (regional utility)
and Solúcar (Abengoa)- prepared the required infrastructure and is responsible for the
operation and maintenance of the system, with support from DLR1 and SBP2.

1
Deutsches Zentrum für Luft- und Raumfahrt e.V. (German Aerospace Center)
2
Schlaich, Bergemann und Partner, GbR.

1
In addition to the Envirodish project objectives, the parabolic dish represents an opportunity
to extend activities of CENTER in the field of solar concentrating systems to this technology,
especially well suited for a new scenario with a growing importance of distributed energy
resources. The development of a detailed thermal model of the engine package of the system,
which will provide a deeper knowledge of the thermal processes, fits into this context.

The ESI Country Reference Unit is connected to the grid and sells the electricity to the local
utility in the frame of the Spanish Special Regime, which rewards electricity generation from
renewable energy resources.

Nomenclature
Parameters and variables
A Heat transfer area (m2)
a Van der Waals constant (J m3/kg2)
b Van der Waals constant (m3/kg)
Cg Parameter in the energy equation..
cp Massic heat capacity at constant pressure (J/kg K)
cv Massic heat capacity at constant volume (J/kg K)
DNI Direct normal Irradiance (W/m2)
F Shape factor
G Mass flow rate from ‘vol’ to ‘vol+1’ (kg/s)
h Coefficient of convective heat transfer (W/m2 K)
J Radiosity
m Mass (kg)
M0 Black body radiation
p Pressure (Pa)
pm Mean pressure (bar)
∆p Pressure drop (Pa)
q Heat transfer flux (W/m2)
Q Heat (J)
Q& Heat rate (kW)
Rg Gas constant for specific gas (J/kg K)
Re Reynolds number (-)
t Time (s)
T Temperature (K)
Tamb Ambient temperature (K)
Tini Initial temperature for gas and regenerator (K)
U& wr Mean heat power stored in the regenerator (kW)
V Volume (m3)
W Work (J)
W& Power (kW)
w Piston speed (m/s)
Ф Crankshaft angle (rad)
Ф0 A fixed crankshaft angle (rad)
∆Ф Crankshaft angle increment (rad)
ε Emissivity
εv Volumetric compression ratio

2
σ Stefan-Boltzmann constant
Subscripts
c Compression
e Expansion
r Regenerator
rad Radiation heat transfer
rc Dead volume between regenerator and compression exchanger
re Dead volume between regenerator and expansion exchanger
r1,r2,…,r10 Regenerator volume 1, volume 2, … , volume 10
vol One of the control volumes
vol+1 The volume at the right of vol in the schema provided bellow
vol-1 The volume at the left of vol in the schema provided bellow
xc Compression exchanger
xc1, xc2 Compression exchanger volume 1, and volume 2
xe Expansion exchanger
xe1, xe2, xe3 Expansion exchanger volume 1, volume 2, and volume 3
w Wall

Methodology
The engine package has been divided in two blocks for modelling purposes: the cavity-
receiver ensemble and the Stirling engine. Both models have been implemented in Matlab,
although the level of detail differs: a simple model for the cavity-receiver ensemble, which
considers radiation heat transfer only, and a detailed engine thermal model of the engine.

Cavity – Receiver Ensemble


A simple radiation transfer model has been developed to estimate radiation losses at the cavity
and receiver. The new ceramic cavity designed for the EnviroDish Project is described in [1].
The geometrical model is a conical frustum with 26 cm base diameter, 12 cm high, and 18.5
cm cover diameter.

The base (surface 1) models the receiver which is the expansion exchanger of the Stirling
engine. Temperatures measured at the receiver have been used to estimate3 T1; the emissivity
of this surface, ε1 = 0.88, has been provided by DLR. The lateral surface (surface 2) is made
of an insulating material, thus it has been imposed that the heat flux is zero. The emissivity of
this surface, ε2 = 0.9, has been provided by DLR4 too.

The cover surface, (surface 3) is an imaginary surface at sky temperature, since most of the
rays leaving the cavity reflect at the concentrator and arrive to the sky. Temperature and
emissivity of this surface have been calculated using [2], being T3=288.0 K and ε3 = 0.87.

3
W. Reinalter. Personal communication. According to this information, there is a difference of 100 ºC approx.
between temperatures at the front side of the receiver and the measured ones since these have been measured
deeper.
4
.W. Reinalter. Personal communication

3
The equations for a surface “i” belonging to a group of n surfaces which contain a closed
volume are available in [3]:
n
J i = ε i ⋅ M i0 + (1 − ε i ) ⋅ ∑ Fij ⋅ J j
j =1
n
q radi = ε i ⋅ M i0 − ε i ⋅ ∑ Fij ⋅ J j
j =1

The equations have been reordered to solve the problem linearly as follows (qrad2=0):

⎛1 − (1 − ε 1 ) ⋅ F11 (1 − ε 1 ) ⋅ F12 (1 − ε 1 ) ⋅ F13 0 ⎞ ⎛ J 1 ⎞ ⎛ ε 1 ⋅ σ ⋅ T14 ⎞


⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜ (1 − ε 2 ) ⋅ F21 1 − (1 − ε 2 ) ⋅ F22 (1 − ε 2 ) ⋅ F23 − ε 2 ⋅σ ⎟ ⎜ J 2 ⎟ ⎜ 0 ⎟
⎜ (1 − ε ) ⋅ F ⎟ ⋅⎜ ⎟ = ⎜ 4⎟
(1 − ε 3 ) ⋅ F32 1 − (1 − ε 3 ) ⋅ F33 0 J ε ⋅ σ ⋅ T3 ⎟
⎜ 3 31
⎟ ⎜ 34 ⎟ ⎜ 3
⎜ −ε ⋅F − ε 2 ⋅ F22 − ε 2 ⋅ F23 ⎟ ⎜ ⎟
ε 2 ⋅ σ ⎠ ⎝ T2 ⎠ ⎜⎝ 0 ⎟
⎝ 2 21 ⎠

Stirling Engine
The Stirling cycle is a thermodynamic regenerative cycle where a gas is alternatively
expanded and compressed in a gas circuit. There are several types of Stirling engines,
described for example in [4] and [5]. The Eurodish Stirling engine is of the α-type. In this
engine two pistons arranged in “V” force the gas to pass back and forth, in an alternative
movement, through the tubes of a heater (the expansion exchanger), a metallic porous mass
(the regenerator), and the tubes of a cooler (the compression exchanger). In the expansion
space there is more mass of gas during the expansion process, while in the compression space
there is more mass during the compression process. A diagram of the engine is presented in
Figure 1

Figure 1. Simplified drawing of the α-type Stirling engine

4
The typical basic Stirling cycle is composed by two constant temperature processes and two
constant volume processes. Practical Stirling engines exhibit a different behavior. According
to the theoretical cycle, there must not be any dead volume and a discontinuity in the piston
displacement is required: None of these requirements is fulfilled by the α-type Stirling engine.
In fact, the displacement of the piston shows a sinusoidal pattern [6].

The Schmidt model, described in [4] and [5], approximates the real behavior of the α-type
Stirling engine by considering a sinusoidal variation of the volumes and allowing for the
existence of dead volumes. This model assumes homogenous pressure in the gas circuit and
constant temperatures on both sides of the regenerator, with a linear temperature distribution
along this element. Therefore, this model does not take into account the tendency of the gas to
cool down during expansion, or to heat up during compression. On the other hand, the
existence of the exchangers is ignored by the model, assuming constant temperatures of the
sources.

The adiabatic model, described by Urieli in [4], is a better approach to the real engine
behavior. In this model the heat transfer occurs in the heat exchangers, while the expansion
and compression volumes behave adiabatically, with variable temperatures over the cycle. A
way to take the engine thermal losses into account is also described in [4].

The model proposed in this work goes further by allowing for variations of mass, pressure and
temperature in the different control volumes; losses, as well as pressure drops, are calculated
taking into account the variable gas temperatures, and the variation of h and the friction factor
with Re.

The gas circuit has been divided in 19 control volumes. Homogenous mass, pressure and
temperature are assumed for each of the control volumes. Ataer and Karabulut [7] analyze the
Stirling cycle dividing the gas circuit in several control volumes with a different approach.

The regenerator mass has been divided in 10 control volumes, according to the control
volumes of the gas into the regenerator, i.e., the gas into volume “r1” is in contact with the
regenerator mass named as “wr1”. The state, mass balance and energy equations are written
for each control volume:

ƒ State equation (Van der Waals):


⎛ a ⋅ m2 ⎞
⎜⎜ p + ⎟⎟ ⋅ (V − m ⋅ b ) = m ⋅ R g ⋅ T
⎝ V2 ⎠

Deriving and ordering the equation in matrix from it results:

5
⎛⎛ ⎞ ⎞ ⎛⎜ dm / dt ⎞⎟
⎜ ⎜ 2 ⋅ a ⋅ m ⋅ (V − b ⋅ m ) − b ⋅ ⎛⎜ p + a ⋅ ⎛⎜ m ⎞⎟ ⎞⎟ − R ⋅ T ⎟
2

(V − m ⋅ b ) (− R g ⋅ m) ⎟ ⋅ ⎜ dp / dt ⎟ =
⎜⎜ V2 ⎜ ⎝ V ⎠ ⎟⎠
g
⎟ ⎟
⎝⎝ ⎝ ⎠ ⎠ ⎜⎝ dT / dt ⎟⎠
⎛ m2 ⎛m⎞ ⎞
2

− ⎜ − 2 ⋅ a ⋅ 3 ⋅ (V − b ⋅ m ) + p + a ⋅ ⎜ ⎟ ⎟ ⋅ (dV / dt )
⎜ V ⎝ V ⎠ ⎟⎠

ƒ Mass balance equation of a generic volume:


− Gvol −1 + (dm / dt ) + Gvol = 0

Where Gvol is the mass flow rate, positive from ‘vol’ to ‘vol+1’; Ge is the mass flow rate from
‘e’ to ‘xe1’; Gxe1 is the mass flow rate from ‘xe1’ to ‘xe2. Derivation and rearrangement gives:

⎛ Gvol −1 ⎞
⎜ ⎟
⎜ dm / dt ⎟
(− 1 1 0 0 1) ⋅ ⎜⎜ dp / dt ⎟⎟ = 0
⎜ dT / dt ⎟
⎜ ⎟
⎝ Gvol ⎠

ƒ Energy equation, in matrix form:


⎛ Gvol −1 ⎞
⎜ ⎟
⎜ dm / dt ⎟
((Cg vol −1 ) (− cv ⋅ T ) 0 (− cv ⋅ m − m ⋅ T ⋅ (dcv / dT )) (− Cg vol )) ⋅ ⎜⎜ dp / dt ⎟⎟ =
⎜ dT / dt ⎟
⎜ G ⎟
⎝ vol ⎠
( p ⋅ (dV / dt ) − h ⋅ A ⋅ (Tw − T ) )
Where:

⎧⎪(c p ⋅ T )vol −1 if Gvol −1 > 0


Cg vol −1 = ⎨
⎪⎩(c p ⋅ T )vol if Gvol −1 < 0

⎧⎪(c p ⋅ T )vol if Gvol > 0


Cg vol = ⎨
⎪⎩(c p ⋅ T )vol +1 if Gvol < 0

ƒ Pressure drops:
p vol − p vol +1 = ∆p ( vol ) − >( vol +1)

Derivation of this equation gives:

(dp vol / dt ) − (dp vol +1 / dt ) = (d∆p ( vol ) −>( vol +1) / dt )

6
The energy equation for each one of the regenerator masses, assuming homogenous
temperature in each regenerator mass, is

d
(mw ⋅ c pw ⋅ Tw ) = h ⋅ Aw ⋅ (T − Tw )
dt

The crankshaft angle vector Ф contains as many revolutions as necessary for the algorithm to
converge. Each revolution can be divided in any number of intervals –800 in this work-
delimited by the points in Ф.

At a given Ф0, start point, pressures, temperatures and masses are known. Then the
derivatives of the state equation, mass equation and energy equations are written for each
control volume resulting in 19·3 = 57 equations. There are also 19-1=18 derived pressure drop
equations which added to the 57 equations referred before make 75 equations. On the other
hand, the mass derivatives, the pressure derivatives, and the temperature derivatives makes
19·3 = 57 unknowns, which added to the 18 mass flow rates make 75 unknowns. The
equations can be solved as a system of linear equations. Once the mass derivatives, the
pressure derivatives, and the temperature derivatives have been calculated, mass, pressure and
temperature in Ф0+∆ Ф are obtained from the definition of derivative.

To keep the linearity of the equations, the pressure drop derivatives have been written at the
right side of the equal sign. However, they cannot be directly calculated since the mass flow
rates are still unknown at Ф0. An iterative process has been used to overcome this obstacle: In
the first iteration, pressure drop is neglected. Mass flow rates are calculated so that they can
be used to calculate ∆p(vol)→(vol+1), which are derived and introduced in the equations to solve
in the next iteration.

The sign of the mass flow rates must be known to write the energy equations. Mass flow rates
are also needed to calculate the coefficients h. But they are unknowns. The following
procedure permits to overcome this obstacle:

ƒ When Ф0=0: coefficients h and the sign of Gvol are set to their inicial values. This is
irrelevant if the initial state is homogenous, since (cp·T)vol = (cp T)vo+1l .
ƒ When Ф0>0: the sign of mass flow rates in Ф0 are set equal to the sign in Ф0-∆Ф Note
that due to continuity, the sign of the mass flow rate will be constant in most calculations,
but an error is made near a mass flow rate change of sign. However in this case the mass
flow rate is near to zero, so the error in the energy equation is negligible.
Each time the calculation in Ф0 is carried out coefficients h and frictions coefficients are
calculated using correlations in [8] and [9]. The energy equations in each one of the ten
masses of the regenerator have to be solved, too. The derivatives of the wall temperatures are
calculated in Ф0 allowing the deduction of the wall mass temperatures in Ф0+∆ Ф through the
derivative definition. Then the wall mass temperatures are used in the next calculation when
Ф0ÆФ0+∆ Ф.

7
Model results
A thorough validation of the model is not possible at this stage, because the measurements
that this validation requires are not yet available. Only the mean pressure at the engine,
receiver temperatures and cooling water temperatures are measured. The mean pressure is
used to determine the hydrogen mass through a short iterative process; temperatures at the
receiver5 and cooling circuit are used to estimate temperatures at the expansion and
compression exchanger, respectively.

To avoid numerical problems due to a high heat transfer rate at the beginning, this
temperatures increase from Tamb in a ramp during NTWXE revolutions, and keep constant
later. This allows a more real simulation by choosing NTWXE according to the time that the
exchangers take to get hot.

Heat transfer coefficients have been calculated using the correlations in [8]; Geometric data
have been taken from the SOLO V-160 information [5], and from drawings in [10] (rough
estimates). Model results for 3 minutes of system operation (May 17th 2005) are presented in
the following. During these 3 minutes, all representative parameters were stable. The
concentrator was clean, thus a reflectivity of 90% has been assumed. Since the effective
concentrator area is 53.099 m2 and the DNI was 773 W/m2, the input power to the cavity is
estimated to be 36.96 kW.

Input data and parameters are presented in Table 1. As a result of the algorithm masses,
temperatures, pressures and mass flow rates are known at any Ф of any cycle. The model
provides transient and steady state results.

Table 1. Input data


Measurements
Variable Value
DNI 773 W/m2
W& 7.55 kW
RPM 1523.8
pm 112.2 bar
Tamb 298.3 K
Relative humidity 30.2%
Model inputs
Variable Value
NV, number of revolutions 50
N, intervals in a revolution 2000
NTWXE, number of cycles to rise the ramp of Twxe 1
NITERATOTAL 2
Mass contained into the engine 0.001715 Kg
Twxe1 951.9 K
Twxe2 1012.8 K
Twxe3 997.2 K
Twxc1 355.4 K
Twxc2 355.4 K
Tini, initial temperature for gas and regenerator 423.2 K

5
See note 1

8
Cavity. The simplified model of the cavity – receiver ensemble gives an estimate of the
radiative losses of this part of the system. The dependence of these losses with receiver
temperature, ambient temperature and relative humidity is shown in Figure 2. Radiative
losses increase with receiver temperature and decrease with ambient temperature and relative
humidity. The radiative losses under the conditions of Table 1 are Q& rad 3 =1.41 kW, with
T2=288.0 K.

Figure 2. From left to right, radiative losses of the cavity-receiver ensemble as a function
of receiver temperature, ambient temperature and relative humidity.

Stirling engine. The model provides transient performance of the engine. Some results for the
last cycle –close to steady state operation- are represented in Figure 3. Temperatures respond
to the expected profile in a Stirling engine: average temperature at the expansion exchanger
wall is higher than Te average. Te average is close to the average temperature of the part of the
regenerator closer to the expansion side (Twr1). The temperature at the regenerator decreases
when moving to the compression side: Twr1>Twr5>Twr10. Tc average value is close to the
average temperature of the regenerator part next to the compression side (Twr10). Finally, the
temperature at the compression exchanger wall is lower than Tc average value.

1100

Twxe2 Twxe3
1000

900 Twxe1 Te
Temperatures (K)

800

Twr5 Twr1
700

600

Twr10
500 Tc

400

Twxc1=Twxc2
300
0 1 2 3 4 5 6

Crankshaft angle (rad). Last revolution

Figure 3. Evolution of selected parameters of the engine versus crankshaft angle for the
last cycle (close to steady state operation).

9
Mean values per cycle of the power exchanges of the Stirling engine have been calculated as
follow:

ƒ Mechanical power delivered:


⎛ ⎞
W& = ⎜⎜ ∫ pe ⋅ dVe + ∫ p c ⋅ dVc ⎟⎟ ⋅ RPM / 60
⎝ revolution revolution ⎠
ƒ Heat power exchange during expansion and compression and thermodynamic energy
increment at the regenerator:
⎛ ⎞
Q& xe = ⎜⎜ ∫ hxe ⋅ ( Axe1 ⋅ (Twxe1 − Txe1 ) + Axe2 ⋅ (Twxe2 − Txe2 ) + Ax3 ⋅ (Twxe3 − Txe3 )) ⋅ dt ⎟⎟ ⋅ RPM / 60
⎝ revolution ⎠
⎛ ⎞
− Q& xc = ⎜⎜ ∫ hxc ⋅ ( Axc1 ⋅ (Txc1 − Twxc1 ) + Axc 2 ⋅ (Txc 2 − Twxc 2 )) ⋅ dt ⎟⎟ ⋅ RPM / 60
⎝ revolution ⎠
&
U wr = Internal Re generator Energy Increment in a cycle ⋅ RPM / 60

ƒ Thermal efficiency
ηth = W& / Q& xe .

The evolution of the mean values per cycle of pressure, temperatures, power exchanges and
efficiency of the Stirling engine is shown in Figure 4. For the last cycle, mean values are:

p c ,mean = 111.3 bar


W& = 11.12 kW
Q& xe = 34.65 kW
U& wr = 0.22 kW
Q& xe ,corrected = Q& xe − U& wr = 34.43 kW
− Q& xc = 23.35 kW
W&
Corrected Efficiency = ⋅ 100 = 32.28%
Q& xe ,corrected

The absorbed heat has been corrected because a fraction of it is still used to heat the
regenerator which must be subtracted. The mean pressure differs slightly from the measured
value. According to Schmidt [ ] mechanical power and heat power exchanged are proportional
to the mean pressure, and so these figures can be scaled:

W& = 11.21 kW
Q& xe ,corrected = 34.72 kW
− Q& xc = 23.54 kW

10
1100

Twxe2 Twxe3
1000

900 Twxe1

Temperatures, mean (K)


800 Te

Twr1
700

600

Twr5
500
Twr10

400
Tc

Twxc1=Twxc2
300
0 10 20 30 40 50 60 70 80 90 100

Revolution

100 35

90 ITERA=1
30
80 ITERA=2
Power magnitudes (kW)

70
25
Efficiency (%)
60

50 20

40 Qxe (kW)
15
30
-Qxc (kW)

20
10
W (kW)
10
Uwr (kW)
0 5
0 10 20 30 40 50 60 70 80 90 100 0 10 20 30 40 50 60 70 80 90 100

Revolution Revolution

Figure 4. Evolution of mean values per cycle of the mean engine pressure (upper left),
temperatures (upper right), power exchanges and thermodynamic energy increment at
the regenerator (lower left) and thermal efficiency of the engine (lower right).
The nominal efficiency of the alternator is 92.5 %. Friction losses have been estimated from
figure 13-9 in [11]. The total friction losses as defined in [11] involve: pumping power, power
to overcome the friction due to relative motion of adjacent components, and the necessary
power to drive the engine accessories. Discounting pumping power (not applicable to Stirling
engines) the estimated friction losses are 0.94 kW.

An exhaustive analysis for a 1 kW-Stirling engine is presented in [12], where the mechanical
losses are estimated as 50% of the engine indicated power. This figure does not seem
applicable for the Solo V-161 due to the different nominal power. The mechanical losses of
the SOLO V-161 must be somewhere in between these two values.

A great uncertainty exists both in the input to the cavity –where reflectivity and spoilage are
estimated- and in the indicated power of the engine, since the real mechanical losses are not
known. At this stage, we can only state that the thermal model of the engine provides detailed
results which are “reasonable” according to the experimental values.

11
Conclusions
The thermal model of the engine package of the Eurodish unit permits a detailed analysis of
both transient and steady state operation of the system. The model of the cavity includes only
radiative losses at this stage, while the Stirling engine model is more detailed and includes
significant improvements over other previous models.

Validation of the model has not been yet possible, since important parameters –power input to
the cavity, temperatures in the engine, shaft power, etc. - are not available, thus leading to
considerable uncertainty. However, the results are reasonable and show a qualitative behavior
similar to what is expected.

Future works will focus on the refinement of the model –both cavity-receiver ensemble and
engine- and validation, this requiring the installation of new sensors at the engine and
measurement of reflectivity and spillage.

References
[1] T. Keck, W. Schiel, P. Heller, W. Reinalter, J-M.Gineste, A. Ferriere, EuroDish –
Continuous Operations, System Improvement and Reference Units. Presented to the
13th International Symposium on Concentrated Solar Power and Chemical Energy
Technologies, Solarpaces, 2006 June 20-23, Seville, Spain.
[2] G. Walton, Thermal Analysis Research program Reference Manual, NBSIR 83-2655,
1983
[3] Grupo de Termotecnia, Dpto. de Ingeniería Energética y Mecánica de Fluidos,
Universidad de Sevilla. Colección de Transparencias de Radiación (October 2000).
[4] http://www.sesusa.org/DrIz/index.html, by I. Urieli.
[5] A. J. Organ. The Regenerator and the Stirling Engine. Mechanical Engineering
Publications Limited, 1997.
[6] J. I. Prieto-García. Funcionamiento del Motor Stirling y Aplicaciones Navales, Doctoral
Thesis, University of Oviedo, 1990.
[7] Ö. E. Ataer, H. Karabulut, Thermodynamic Analysis of the V-type Stirling-Cycle
Refrigerator. International Journal of Refrigeration 28 (2005) 183-189
[8] Grupo de Termotecnia, Dpto. de Ingeniería Energética y Mecánica de Fluidos,
Universidad de Sevilla. Colección de Tablas y Gráficas de Termotecnia, version 3.0
(September 2002)
[9] W. Kays, A. L. London, Compact Heat Exchangers 2nd edition, McGraw-Hill,1964.
[10] Service and Operator’s Manual for the SOLO 161 Solar Stirling Unit. Solo Stirling,
March 2004.
[11] J. B. Heywood, Internal Combustion Engine Fundamentals, New York, McGraw-Hill,
1988.

12
[12] K. Makhkamov and D.B. Ingham Determination of the Performance of a Stirling Engine
as used in a Dish/Stirling System. European Stirling Forum 1998, Osnabruck, Germany,
24-26 February 1998.

Acknowledgements
The authors wish to thank Lotta Koch and Alexander Wilhelm Kromp, Erasmus students at
the Seville Engineering School, for their collaboration in the analysis of the heat transfer at
the cavity, and Wolfgang Reinalter (DLR) for his support and for providing relevant
information.

13

You might also like